Vertical Electronic Excitations in Solution with the EOM-CCSD Method

May 29, 2013 - The accurate calculation of electronic transition energies and properties of isolated chromophores is not sufficient to provide a reali...
0 downloads 0 Views 1002KB Size
Subscriber access provided by University of Glasgow Library

Article

Vertical Electronic Excitations in Solution with the EOM-CCSD Method Combined with a Polarizable Explicit/Implicit Solvent Model Marco Caricato, Filippo Lipparini, Giovanni Scalmani, Chiara Cappelli, and Vincenzo Barone J. Chem. Theory Comput., Just Accepted Manuscript • DOI: 10.1021/ct4003288 • Publication Date (Web): 29 May 2013 Downloaded from http://pubs.acs.org on June 3, 2013

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Chemical Theory and Computation is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Vertical Electronic Excitations in Solution with the EOM-CCSD Method Combined with a Polarizable Explicit/Implicit Solvent Model Marco Caricato,∗,† Filippo Lipparini,‡ Giovanni Scalmani,† Chiara Cappelli,‡,¶ and Vincenzo Barone‡ Gaussian, Inc., 340 Quinnipiac St. Bldg. 40, Wallingford, Connecticut 06492, USA., and Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy. E-mail: [email protected]

Abstract The accurate calculation of electronic transition energies and properties of isolated chromophores is not sufficient to provide a realistic simulation of their excited states in solution. In fact, the solvent influences the solute geometry, electronic structure, and response to external fields. Therefore, a proper description of the solvent effect is fundamental. This can be achieved by combining polarizable explicit and implicit representations of the solvent. The former provides a realistic description of solvent molecules around the solute, while the latter introduces the electrostatic effect of the bulk solution and reduces the need of too large a number of explicit solvent molecules. This strategy is particularly appealing when an accurate method such as equation of motion coupled cluster singles and doubles (EOM-CCSD) ∗ To

whom correspondence should be addressed Gaussian, Inc., 340 Quinnipiac St. Bldg. 40, Wallingford, Connecticut 06492, USA. ‡ Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy. ¶ Dipartimento di Chimica e Chimica Industriale, Università di Pisa, via Risorgimento, 35 I-56126 Pisa, Italy



1 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

is employed for the treatment of the chromophore. In this contribution, we present the coupling of EOM-CCSD with a fluctuating charges (FQ) model and polarizable continuum model (PCM) of solvation for vertical excitations in a state-specific framework. The theory, implementation, and prototypical applications of the method are presented. Numerical tests on small solute-water clusters show very good agreement between full EOM-CCSD and EOM-CCSDFQ calculations, with and without PCM, with differences ≤ 0.1 eV. Additionally, approximated schemes that further reduce the computational cost of the method are introduced and showed to perform well compared to the full method (errors ≤ 0.1 eV).

2 ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

1 Introduction The accurate simulation of photochemical processes has become increasingly important in the past decade as an invaluable tool to assist and often guide experiments. The study of electronic excited states is at the frontier of modern science and goes beyond traditional chemistry to include multidisciplinary research fields, as in energy and material science. This kind of applications, however, requires a realistic description of the complex environment around the chromophore because of the important effects on its electronic structure. The computational resources required for treating large systems are still much larger than what is achievable with standard quantum mechanical (QM) tools especially when accurate methods are employed as, for instance, the series of methods provided by coupled cluster theory (CC) 1,2 . A reasonable compromise between accuracy and computational effort can be achieved by resorting to multiscale or “focused” models, where the system is separated in multiple layers and the target structure is treated with a higher level of theory while the rest is described with less accurate but computationally less demanding approaches. The prototypical “environment” surrounding a chromophore is a solvent. Much effort has been devoted to reliable and yet affordable modeling of the large number of molecules required to properly describe solution phase. In general, solvation models can be separated in two large categories: explicit and implicit models. The former maintain an atomistic representation of the solvent, thus are able to describe specific solute-solvent interactions (e.g., hydrogen bonds) and provide a more realistic picture. The drawback is that explicit models require a large number of solvent molecules to properly describe bulk effects and, consequently, require the consideration of a large number of configurations in order to achieve a statistical average of a given property. Implicit models, on the other hand, replace the atomistic representations of the solvent with a continuum, polarizable medium. The advantage here is a lower computational cost since long-range bulk effects and statistical averaging are naturally accounted for. The drawback is that specific solute-solvent interactions cannot be properly described. A large number of models have been developed in both families and a comprehensive account of all of them is beyond the scope of this work 3–50 . The best of the two worlds can be obtained by a judicious combination of the two approaches, 3 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

which would limit the number of explicit solvent molecules to the first few solvation shells (to account for specific solute-solvent interactions) while including long-range effects through a continuum model. In this way, the number of explicit solvent configurations (and corresponding QM calculations) necessary for statistical averaging can be greatly reduced. This strategy assumes critical importance when a member of the CC family of methods is used, such as the equation of motion singles and doubles (EOM-CCSD) 1,2,51–54 . In recent publications, Lipparini et al. 19–21 reported the combination of a fluctuating charge (FQ) 16–18 model and the polarizable continuum model (PCM) 22–25,27 for ground and excited states while Caricato 49,50 reported the combination of EOM-CCSD with PCM. In this contribution, we present the combination of these approaches in the EOM-CCSD-FQ and EOM-CCSD-FQ-PCM methods, and their implementation. The formulation of FQ and FQ-PCM in Ref. 19,20, whose equations are formally very similar to those for PCM, allows for a relatively direct extension of the EOM-CCSD-PCM method to this explicit/implicit approach, including approximate schemes that can be used to further reduce computational cost 49,50 . The test systems presented in this work are small solute-water clusters (including up to six explicit water molecules). These are certainly not enough to provide a realistic representation of a solvation shell but can be treated entirely at EOM-CCSD level, thus providing a direct comparison between full QM and hybrid calculations. Larger, more realistic solvation shells will be the subject of future investigation. The results in this paper show the efficacy of this focused model for the reliable computation of solvatochromic shifts compared to full QM calculations and open new avenues for the application of accurate QM methods to simulation of real-life systems. This paper is organized as follows. The theory and implementation of the FQ-PCM models and their coupling with EOM-CCSD is presented in Section 2. Numerical tests are reported in Section 3, and a discussion of the results and concluding remarks are presented in Section 4.

4 ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

2 Theory In this section, we review the theory for the FQ model combined with PCM and their coupling with the potent EOM-CCSD method for electronic excited states. Thanks to the formulation of the FQ and FQ-PCM models in Ref. 19,20, the coupling with EOM-CCSD straightforwardly follows that with PCM 49,50 , as discussed in Section 2.2. Therefore, here we only report the main expressions and refer the reader to Refs. 19,20,49,50 for details on their derivation and implementation.

2.1 FQ-PCM In the FQ model, the polarizability of the force field is introduced by endowing each atom with a point charge, whose value depends on the environment 16 according to the electronegativity equalization principle (EEP). 55,56 The EEP can be recast in a variational formulation, which is a convenient way to define FQs as the ones minimizing the following functional: F(Q, λ ) = ∑ Qα i χα i + α ,i

1 ∑ Qα iJα i,β j Qβ j + ∑ λα (∑ Qα i − Qtot α ) 2∑ α i α ,i β , j

1 =Q · χ + Q · JQ + λ · Q 2

(1)

where the Greek indexes α and β run on molecules and the Latin ones on the atoms of each molecule. In Eq. (1), the χ vector contains the electronegativities of the atoms, the matrix J represents the (screened) electrostatic interaction between the fluctuating charges Q, and a set of Lagrangian multipliers λα is used to impose charge conservation constraints. By introducing a compact notation (see Ref. 19 for a detailed derivation and description of the various parameters that define the model), the stationarity conditions read

DQλ = −CQ

5 ACS Paragon Plus Environment

(2)

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

where CQ is a vector containing atomic electronegativities and total charge constraints, D is the matrix composed by J and the Lagrange multipliers related blocks, whereas Qλ is a vector containing charges and Lagrange multipliers. Consistently with the semi-classical nature of a QM/MM approach, we introduce the coupling with a QM (SCF) description by treating the interaction between the two densities of charge (i.e., the one represented by the FQs and the QM one) classically, i.e., through a simple Coulomb interaction: Nq

EQM/MM = ∑ Φ[ρQM ](ri )Qi

(3)

i=1

where the sum runs over the Nq FQs and Φ[ρQM ](ri ) is the electrostatic potential due to the QM density of charge at the i-th FQ placed at ri : de f

Φ[ρQM ](ri ) =

N e V i (P) = V i (P) +V i (P) =

Z Zζ ρ el (r) − dr ∑ |r − Rζ | R3 |ri − r| ζ =1 i Nn

Here, the ζ -labeled sum runs over the Nn QM nuclei, whose position we call Rζ (see Ref. 20,21 for a detailed discussion of the coupling). For later convenience, we have introduced a slightly modified notation, indicating with and overbar the expectation values of the electrostatic potential and of the charges to distinguish them by the associated operators. This distinction can be made clearer for the second term of the potential by expanding the electronic density ρ el (r) in terms of an atomic basis set: V i (P) = − ∑ Pµν e

µν

Z

R3

dr

χµ (r)χν (r) = ∑ Pµν Vµν ,i |ri − r| µν

(4)

where the “uncontracted” potential Vµν , which is the representation of the potential operator in the atomic orbitals (AO) basis set, is introduced and the double sum runs on the AO: the expectation value is obtained by contracting the operator with the density matrix. The definition of the interaction adopted yields a QM/MM energy functional that is variational with respect to both densities of charge, once the suitable constraints have been taken into account,

6 ACS Paragon Plus Environment

Page 7 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

in a similar fashion to what has been proposed by some of the present authors for PCM 35,57 . Therefore: 1 1 E [P, Q, λ ] = trhP + trPG(P) + Q · χ + Q · JQ + λ · Q + Q · V(P) 2 2

(5)

where h and G are the usual one- and two-electron matrices. It is possible to introduce the coupled Fock/FQ equations by introducing an extended Fock operator, which can be computed as the gradient of the energy functional Eq. (5) with respect to the density matrix:

∂E F˜ = = h + G(P) + Q · V ∂P

(6)

The FQs are obtained by imposing the stationarity of the global functional with respect to the charges and the Lagrangian multipliers; with respect to Eq. (2), a new source term appears:

DQ = −C − V(P)

(7)

Such term represents the coupling between the QM and MM portions of the system. The inclusion of a further layer treated by means of the PCM approach (in particular, we choose to the conductorlike PCM 58–61 ) is easily obtained by redefining the charges and the D matrix so as to also include PCM contributions: 



FQ









FQ



Ω  Q   C   V (P)   D  −  PCM   PCM  = −  .  Ω S/ f (ε ) 0 Q V (P) †

(8)

where S and Ω represent the Coulomb interaction of the PCM charges with themselves and with the FQs, respectively, and we have introduced the superscripts FQ and PCM to distinguish the two PCM

sets of charges. In addition, a PCM contribution Q

FQ

· V is to be added to Q

· V in Eq. (6).

Further details on the coupling of the PCM with the FQ model can be found elsewhere. 19,20

7 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 25

2.2 EOM-CCSD-FQ-PCM The coupling of CCSD for ground and excited states with the FQ model directly follows its coupling with PCM 44–50 . The complete scheme is called PTED (perturbation theory energy and density) for historical reasons 62,63 , and the solvent reaction field depends on the correlation density (thus the D in the acronym). For the ground state, a CC free energy functional can be written as: E 1 D ξ ξ −T ξ T ED + Φ (1 + Λ)e H GPT = G e 0 0 N Φ0 + VN · QN re f 2

(9)

where Φ0 is the reference wave function (HF in most cases), Gre f is the associate reference solvation free energy, T and Λ are the CC excitation and de-excitation operators. The subscript N ξ

indicates the normal-product form of an operator (XN = X − hΦ0 |X | Φ0 i), and the Hamiltonian HN includes the reference charges: ξ

ξ

HN = HN + VN · Q0 .

(10)

The superscript ξ refers to the solvation model in use: FQ or FQ-PCM. For FQ-only, the refFQ

erence charges Q0 , which depend on the nuclei and SCF electronic charge density, are computed FQ

by solving Eq. (7). The charges QN , on the other hand, depend on the correlation density:

γ pq = hΦ0 | (1 + Λ) e−T {p† q}eT |Φ0 i .

(11) FQ

where p, q are two generic molecular orbitals (MOs). The charges QN are then obtained by solving: FQ

DQN = −VN (γ ) .

(12)

For the combined FQ-PCM approach, the reference charges are computed by solving Eq. (8) while the correlation charges by solving: 



FQ QN





FQ VN ( γ )



Ω     D   PCM  = −  PCM  .  Ω S/ f (ε ) QN VN ( γ ) †

8 ACS Paragon Plus Environment

(13)

Page 9 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

In this case, the dimension of the charges vector in Eq. (9) is the sum of the number of FQ and PCM charges. From the point of view of the CC equations (and associated computer code), however, the choice of solvation model is transparent and the same machinery introduced for CC-PCM can be employed 44,46,47 . In this work, we have developed the interface of the code that computes the FQ and FQ-PCM charges in Eqs. (12) and (13) with the code that produces the corresponding operators that enter the CC equations for the evaluation of Eq. (9). It is important to note that the computational cost for the evaluation of the charges in Eqs. (7), (8), (12) and (13) is negligible compared to the solution of the CC equations. ED is computed by minimizing the expression in Eq. (9) with respect to the T and Λ amGPT 0

plitudes. The computational demand is thus at least twice as large as that for the corresponding gas phase calculation. Similar to PCM, approximations can be introduced to avoid the need of solving for the Λ amplitudes. These approximate schemes, called PTE, PTE(S), and PTES do not require the calculation of the entire correlation density but only the CC amplitudes. In PTE, the explicit solvation term in the correlation part (last term in Eq. (9)) is completely neglected, and the solvent effect is only introduced in the reference wave function through an explicit energy term and implicit orbital polarization. In PTE(S) and PTES, an explicit solvation term is introduced in the correlation equations but is computed by approximating the correlation density with the only term that does not depend on the Λ amplitudes, i.e., the T singles amplitudes (thus the “S” in the name). In PTE(S), such term is only a correction to the PTE free energy while in PTES there is an explicit operator in the T equations. Details for these schemes can be found in Refs. 46,47,49,50. Similarly to ground state, the K-th excited state free energy functional takes the form: i h ξ −TK ξ TK ED e , R |Φ0 i + hΦ | L e H GPT =G K K 0 K N re f E D ξ + Φ0 (1 + ΛK )e−TK HN eTK Φ0

(14)

1 K K,ξ + ωK (1 − hΦ0 | LK RK |Φ0 i) + VN · QN 2

where ωK , LK , and RK are the EOM-CC similarity tranformed Hamiltonian K-th eigenvalue, left,

9 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

K,ξ

and right eigenvectors, respectively. The charges QN are computed by solving Eq. (12) or Eq. (13) for FQ or FQ-PCM approaches, respectively, with the K-th state density: h i K = hΦ0 | LK e−T {p† q}eT , RK |Φ0 i γ pq + hΦ0 | (1 + ΛK ) e

−T



(15)

T

{p q}e |Φ0 i .

ED is obtained after minimization of the functional in Eq. (14) with respect to the T , Λ , GPT K K K

LK , and RK amplitudes. Since this is a state-specific approach, this set of amplitudes is different for each excited state. In order to avoid the computation of the ΛK amplitudes and decouple the ground state from the excited state part of the calculation, the PTE, PTE(S), and PTES approximate schemes can be introduced also for EOM-CC. For excited states, the PTE scheme includes statespecific solvent terms in the EOM equations. Therefore, the scheme where no explicit term appears in the CC equation is called FRF, for frozen-reaction-field (FRF and PTE are the same for the ground state). More details can be found in Ref. 49,50. When vertical transition calculations are performed with a continuum solvation model, the different response-time of the electrons and nuclei of the solvent molecules must be taken into account. This is achieved in PCM by introducing a nonequilibrium solvation regime 25,64,65 , which represents the situation where only the solvent electron response is allowed to equilibrate with the new solute electronic density while the solvent nuclei response is kept frozen in the initial state. This approach is not necessary for the FQ method alone since a nonequilibrium response is automatically implied as the solvent molecules are not allowed to move. An explicit nonequilibrium calculation is, on the other hand, necessary for the FQ-PCM approach. The solvent electrons response is obtained by replacing the equilibrium dielectric constant ε in Eq. (13) with the optical dielectric constant ε∞ . The nonequilibrium EOM-CCSD-FQ-PCM expressions for the PTED, PTES, PTE(S), and PTE schemes are identical to those for EOM-CCSD-PCM as reported in Ref. K,FQ−PCM

50, provided that the charges QN

in Eq. (13) are computed using ε∞ .

10 ACS Paragon Plus Environment

Page 11 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

3 Results The methods described in the previous section, for the complete and approximate schemes, are tested here on three molecules with small water clusters: formaldehyde, p-nitroaniline (PNA), and pyrimidine. The first two molecules are considered with two, four and six water molecules, and are shown in figure Figures 1 and 2 where the “solute” molecules are depicted as balls and sticks while the water molecules as tubes. The geometries are taken from Ref. 12 where EOM-CCSD was combined with the effective fragment potential (EFP) 8,9 method. This allows for a direct comparison of the two approaches, FQ and EFP. Although, it is important to mention some differences between them: FQ is an electrostatic-only model but includes mutual solute-solvent polarization effects in the CC equations due to the iterative solution of Eq. (9) and Eq. (14); EFP in Ref. 12, on the other hand, includes other effects at the SCF level (i.e.,dispersion, exchange-repulsion, and polarization) but only a polarization correction to the excitation energy after conversion of the EOM equations. We also consider the effect of a continuum solvation model around the entire cluster. The latter molecule, pyrimidine, is considered with two water molecule around it in such a way that the cluster structure maintains C2v symmetry, shown in Figure 3. The geometry of this cluster is taken from Ref. 20. In general, the comparison will be made between the full EOM-CCSD calculation for the solute-water clusters and that with the EOM-CCSD-FQ approach, both with and without PCM. The 6-31+G(d) basis set is used for the calculations on formaldehyde and PNA, as in Ref. 12. For formaldehyde, also aug-cc-pVDZ 66 is used to assess some basis set effects. For pyrimidine, the aug-N07 basis set 67 is used as in Ref. 20. The core orbitals are kept frozen in the CC part of the calculation except for formaldehyde with the 6-31+G(d) basis set. The PCM cavity is built with a series of interlocking spheres centered on the nuclei and employing solvation model D (SMD) radii 68 . The CPCM 58–61 flavor of PCM is used for all calculations. The FQ parameters for water are ηO = 367.0 kcal/mol2, ηH = 392.2 kcal/mol2, and χOH = 73.33 kcal/mol. These parameters are taken directly from Ref. 16, and no optimization was attempted. All calculations are performed with a development version of the GAUSSIAN suite of programs 69 . 11 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 25

3.1 Formaldehyde Figure 1: Structure of the formaldehyde-water clusters.

(a)

(b)

(c)

Table 1: Formaldehyde excitation energy (eV) for the n → π ∗ transition with the 6-31+G(d) basis set with two, four, and six water molecules. “Full” refers to the calculation done on the entire system with the EOM-CCSD method. The EFP results are from Ref. 12. The excitation energy for the isolated molecule is 4.137 eV. Full FRF PTE PTE(S) PTES PTED EFP

2 4.283 4.278 4.257 4.255 4.255 4.261 4.334

4 4.240 4.246 4.231 4.228 4.229 4.234 4.286

6 4.421 4.337 4.313 4.310 4.311 4.318 4.430

The n → π ∗ transition energy for the isolated formaldehyde is 4.137 eV with the 6-31+G(d) basis set. The excitation energies for the formaldehyde-water clusters in gas phase computed with the same basis set are shown in Table 1. The “Full” results refer to the calculations computed entirely at EOM-CCSD level, while the following rows report the results obtained by considering 12 ACS Paragon Plus Environment

Page 13 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Table 2: Formaldehyde excitation energy (eV) for the n → π ∗ transition with the 6-31+G(d) basis set with two, four, and six water molecules and PCM. “Full-PCM” refers to the calculation done on the entire system with the EOM-CCSD-PCM schemes. 2 FRF PTE PTE(S) PTES PTED

4

Full-PCM FQ-PCM 4.436 4.472 4.351 4.362 4.319 4.327 4.330 4.339 4.374 4.360

Full-PCM 4.410 4.325 4.292 4.303 4.349

6 FQ-PCM 4.451 4.344 4.310 4.322 4.342

Full-PCM 4.510 4.431 4.402 4.413 4.457

FQ-PCM 4.479 4.373 4.340 4.352 4.373

the water molecules as classical, polarizable charges according to the various schemes described in Section 2.2. Finally, the last row reports the results using the EFP method from Ref. 12. The agreement between the EOM-CCSD-FQ method and the full EOM-CCSD calculations is very good: 0.02 eV, 0.01 eV, and 0.10 eV for the 2-, 4-, and 6-water clusters, respectively, with the complete PTED scheme. The performance of FQ and EFP is similar, with the former method being closer to the reference data for the two smaller clusters, and the latter being better for the larger cluster. A comparison between the various approximate schemes for the EOM-CCSD-FQ approach indicates differences not larger than 0.02 eV. This is important from a computational point of view since the complete PTED scheme requires two separate calculations, for the ground and excited states, thus it is roughly twice as expensive as the approximate schemes. Even the FRF scheme (where the FQ charges are frozen at their SCF values) performs well, although, as we shall see when the PCM contribution is added, this approximation is in general too drastic. Among the approximate schemes, PTE is the closest to PTED, consistently with Ref. 49,50. Introducing bulk solvation effects through PCM shifts the excitation energies of the clusters considerably. The data is reported in Table 2. “Full-PCM” refers to the full QM calculations while “FQ-PCM” indicates the QM-classical approach. Since various schemes can be also defined for the full EOM-CCSD-PCM calculations, the comparison between the Full-PCM and FQ-PCM results can be directly carried out for all schemes. First, however, notice that the PCM effect on the full PTED calculation shifts the transition energy by 0.09 eV, 0.11 eV, and 0.04 eV for the 2-, 4-, and 6-water clusters, respectively. This is a indication that these small clusters cannot be considered a 13 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

Table 3: Formaldehyde excitation energy (eV) for the n → π ∗ transition with the aug-cc-pVDZ basis set with two, four, and six water molecules. “Full” refers to the calculation done on the entire system with the EOM-CCSD method. The excitation energy for the isolated molecule is 4.125 eV.

Full FRF PTE PTE(S) PTES PTED

2 4.276 4.273 4.251 4.248 4.248 4.254

4 4.227 4.239 4.222 4.219 4.220 4.225

6 4.410 4.334 4.309 4.306 4.307 4.314

Table 4: Formaldehyde excitation energy (eV) for the n → π ∗ transition with the aug-cc-pVDZ basis set with two, four, and six water molecules and PCM. “Full-PCM” refers to the calculation done on the entire system with the EOM-CCSD-PCM schemes.

FRF PTE PTE(S) PTES PTED

2 Full-PCM FQ-PCM 4.436 4.472 4.348 4.358 4.317 4.323 4.328 4.335 4.374 4.356

4 Full-PCM 4.408 4.319 4.287 4.298 4.347

6 FQ-PCM 4.451 4.341 4.306 4.318 4.338

Full-PCM 4.503 4.422 4.395 4.405 4.451

FQ-PCM 4.479 4.369 4.335 4.347 4.368

realistic representation of a solvated system. However, since our goal is to compare with full-EOM calculations, we are limited in the system size we can treat. The agreement between the Full-PCM and FQ-PCM is very good: of the order of 0.02 eV in most cases; the largest difference is found for the 6-water cluster with the PTED scheme, 0.08 eV, which is smaller than the same results in gas phase (0.10 eV, see Table 1). The difference between the approximate schemes and PTED is small 0.02-0.04 eV, except for the FRF scheme where the difference can be > 0.1 eV. The best agreement with PTED is achieved with the PTE scheme. Changing basis set, from 6-31+G(d) to aug-cc-pVDZ, does not significantly change the transition energy. The data with the latter basis is reported in Table 3 for the gas phase clusters. The transition energy for the isolated molecule becomes 4.125 eV, thus a little more than 0.01 eV lower than with the Pople basis set. Similar changes are obtained for the full QM cluster calculations. The FQ calculations also follow similar trends, so the results are consistent with those discussed

14 ACS Paragon Plus Environment

Page 15 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

above for the 6-31+G(d) basis. The effect of PCM, shown in Table 4, is to reduce even further the difference between basis sets. Overall, the best agreement with PTED is again obtained by employing the PTE approximate scheme.

3.2 PNA Figure 2: Structure of the PNA-water clusters.

(a)

(b)

(c)

Table 5: PNA excitation energy (eV) for the π → π ∗ transition with the 6-31+G(d) basis set with two, four, and six water molecules. “Full” refers to the calculation done on the entire system with the EOM-CCSD method. The EFP results are from Ref. 12. The excitation energy for the isolated molecule is 4.654 eV. Full FRF PTE PTE(S) PTES PTED EFP

2 4.375 4.490 4.479 4.480 4.480 4.482 4.384

4 4.461 4.536 4.525 4.527 4.527 4.529 4.459

6 4.391 4.511 4.490 4.492 4.492 4.497 4.415

The π → π ∗ transition energy for the PNA-water clusters in gas phase is reported in Table 5. The difference between the FQ results with the PTED scheme and the full EOM calculations is 0.11 eV, 0.07 eV, and 0.11 eV for the 2-, 4-, and 6-water cluster, respectively, which is slightly worse than the EFP results but still much smaller than the shift compared to isolated PNA (0.28 eV, 0.19, and 0.26 eV, respectively). All the approximate schemes perform very well, with differences below 0.01 eV from PTED. 15 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 25

Table 6: PNA excitation energy (eV) for the π → π ∗ transition with the 6-31+G(d) basis set with two, four, and six water molecules and PCM. “Full-PCM” refers to the calculation done on the entire system with the EOM-CCSD-PCM schemes. 2 FRF PTE PTE(S) PTES PTED

4

Full-PCM FQ-PCM 3.902 3.908 3.612 3.598 3.650 3.639 3.650 3.639 3.743 3.779

Full-PCM 3.916 3.625 3.664 3.664 3.757

6 FQ-PCM 3.915 3.605 3.646 3.646 3.786

Full-PCM 3.929 3.645 3.684 3.684 3.775

FQ-PCM 3.922 3.607 3.648 3.648 3.789

The introduction of bulk solvation effects shifts the excitation energy to much lower values as shown in Table 6 when the effect of PCM is considered. This red-shift of about 0.9 eV (for all clusters) indicates a strong stabilization of the excited state energy compared to the ground state. This is due to a charge separation in the excited state that is stabilized by the presence of a polar solvent like water. The difference between the energy of various clusters is now within 0.01 eV for all Full-PCM calculations with the PTED scheme, contrary to differences of more 0.1 eV for the corresponding results in gas phase (c.f. Table 5). The difference between the various schemes for the Full-PCM calculations is slightly larger than compared to gas phase. The PTES results are less than 0.1 eV from the complete PTED scheme, while the FRF scheme provides the worst performance with difference above 0.15 eV. More importantly, the comparison between the Full-PCM and FQ-PCM results shows very good agreement, with differences below 0.04 eV.

3.3 Pyrimidine Figure 3: Structure of the pyrimidine-water cluster.

16 ACS Paragon Plus Environment

Page 17 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Table 7: Pyridine excitation energy (eV) for the n → π ∗ (B1 ) transition with the aug-N07 basis set. “P” refers to pyrimidine alone while “P+2W” indicate the cluster with two water molecules. The transition energy computed with full EOM-CCSD on the isolate molecule and the water cluster without PCM are 4.581 eV and 4.899 eV, respectively. FRF PTE PTE(S) PTES PTED

P/PCM 4.814 4.741 4.735 4.737 4.771

P+2W/FQ 4.908 4.864 4.862 4.862 4.870

P+2W/PCM 5.015 4.952 4.949 4.950 4.982

P+2W/FQ-PCM 5.068 4.962 4.956 4.957 4.990

The summary of the results for the pyrimidine-water cluster is reported in Table 7. The transition considered is the lowest n → π ∗ (B1 symmetry). The transition energy for the isolated molecule and the cluster in gas phase for the full QM calculation is 4.581 eV and 4.899 eV, respectively. Therefore, the two water molecules produce a shift of 0.32 eV. This can be compared with the shift obtained with PCM on pyrimidine alone (see the “P/PCM” column in Table 7), which is 0.19 eV. Note that the approximate schemes in this case are within 0.04 eV from PTED. The comparison between the full EOM calculation for the water cluster in gas phase and that with the FQ schemes indicates a very good performance of the FQ approach (see the “P+2W/FQ” column in Table 7): the difference of the PTED scheme from the full EOM result is below 0.03 eV, while the difference between the approximate schemes and PTED is below 0.01 eV (except for FRF). The addition of PCM on top of the pyrimidine-water cluster blue shifts the transition energy of about 0.1 eV (see the “P+2W/FQ-PCM” column in Table 7). The comparison between the full EOM-PCM and the EOM-FQ-PCM calculation is very positive, with differences below 0.01 eV. For this system, PTE is the closest approximate scheme to PTED.

4 Discussion and Conclusions This work reports the coupling of the highly accurate EOM-CCSD method for excited states with a combination of explicit/implicit polarizable solvent models: FQ and PCM. This approach aims at introducing the essential solvent effect on excitation energies when these are computed with

17 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

computationally demanding methods such as those offered by CC theory while maintaining a reasonable balance between accuracy and computational cost. In this direction, the combination of the FQ and PCM approaches is a decisive step towards a reliable description of direct solute-solvent interaction and bulk solvent effects. Indeed, the use of a continuum model like PCM has the potential advantage of reducing the number of explicit solvent molecules necessary to obtain convergence on long-range interactions, which otherwise can only be reproduced by considering a large number of solvent configurations (provided, for instance, by a molecular dynamics simulation). Reducing the number of configurations is extremely important when the solute is treated at CC level because it reduces the number of EOM-CC calculations. Although our goal is to use the EOM-CCSD-FQ-PCM method for realistic modeling of solvated systems, which still implies a relatively large number of explicit solvent molecules, the examples reported in this paper focus on small solute-water clusters, with and without PCM. In this way, the entire cluster can be treated at EOM-CCSD level thus allowing for a direct comparison with the full QM calculation. The first two systems, formaldehyde and PNA (see Section 3.1 and Section 3.2), were used to test the popular EFP approach 12 and thus offer an opportunity for a comparison with a different solvation model. Note, however, that our approach includes mutualpolarization between solute and solvent, which are reported only as an a posteriori correction in Ref. 12. Although such effects are small for transition energies, they can be significant for excited state properties 49 . The performance of FQ for the gas phase clusters is rather good, with differences ≤ 0.1 eV from the full QM calculations and comparable to EFP. Further improvements in our method may be obtained by including other effects, like dispersion and exchange-repulsion, already present in EFP. Also, a better parameterization of the model for post-SCF methods would be beneficial. Since Lipparini et al. 19–21 casted the FQ expressions in a way formally similar to those of PCM, the coupling of the two models is straightforward, as summarized in Section 2.1. This has also the advantage that the (EOM-)CCSD-PCM series of methods 44–50 can be seamlessly extended to the coupling with FQ and FQ-PCM. When the EOM-CCSD-FQ-PCM results are compared with

18 ACS Paragon Plus Environment

Page 19 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

those from full EOM-CCSD-PCM calculations, the agreement is even better than in gas phase, see Table 2, Table 4, and Table 6. The results reported in Section 3.3 for the last test system, a cluster of pyrimidine and two water molecule in C2v configuration, show an even better performance of both the FQ and FQ-PCM approaches compared to the full QM calculations, with differences below 0.04 eV and 0.01 eV, respectively. The approximate schemes perform all quite well. The FRF scheme, which does not include mutual-polarization effects, provides errors larger than 0.1 eV (and larger than 0.15 eV for PNA, see Table 6) and is, as expected, the least reliable. The appeal of the approximate schemes is that the transition energy can be computed in one step, while PTED requires two separate calculations for the ground and the excited states. In summary, the results presented in this work hold the promise of reliable and yet computationally feasible calculations of excited states of chromophores in solution by judiciously coupling a highly accurate method for the solute (EOM-CCSD) and an approximate polarizable explicit/implicit model (FQ-PCM) for the solvent. Simulations of realistic solvated systems are currently in progress in our laboratories.

5 Acknowledgments VB acknowledges the European Research Council (ERC) for financial support through the Advanced Grant DREAMS: 320951 “Development of a Research Environment for Advanced Modeling of Soft matter”. CC thanks the Italian MIUR (PRIN 2009: Sviluppo di modelli accurati e di codici veloci per il calcolo di spettri vibrazionali and FIRB-Futuro in Ricerca Protocollo: RBFR10Y5VW) and COST (Action CODECS: “COnvergent Distributed Environment for Computational Spectroscopy”) for financial support.

References (1) Bartlett, R. J.; Musial, M. Rev. Mod. Phys. 2007, 79, 291–352.

19 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

(2) Shavitt, I.; Bartlett, R. J. Many-Body Methods in Chemistry and Physics; Cambridge University Press: Cambridge, 2009. (3) Osted, A.; Kongsted, J.; Mikkelsen, K. V.; Christiansen, O. J. Phys. Chem. A 2004, 108, 8646–8658. (4) Kongsted, J.; Osted, A.; Pedersen, T. B.; Mikkelsen, K. V.; Christiansen, O. J. Phys. Chem. A 2004, 108, 8624–8632. (5) Aidas, K.; Kongsted, J.; Osted, A.; Mikkelsen, K. V.; Christiansen, O. J. Phys. Chem. A 2005, 109, 8001–8010. (6) Kowalski, K.; Valiev, M. J. Phys. Chem. A 2006, 110, 13106–13111. (7) Fan, P.-D.; Valiev, M.; Kowalski, K. Chem. Phys. Lett. 2008, 458, 205–209. (8) Day, P. N.; Jensen, J. H.; Gordon, M. S.; Webb, S. P.; Stevens, W. J.; Krauss, M.; Garmer, D.; Basch, H.; Cohen, D. J. Chem. Phys. 1996, 105, 1968–1986. (9) Gordon, M. S.; Freitag, M. A.; Bandyopadhyay, P.; Jensen, J. H.; Kairys, V.; Stevens, W. J. J. Phys. Chem. A 2001, 105, 293–307. (10) Bandyopadhyay, P.; Gordon, M. S.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 2002, 116, 5023–5032. (11) Sneskov, K.; Schwabe, T.; Kongsted, J.; Christiansen, O. J. Chem. Phys. 2011, 134, 104108. (12) Slipchenko, L. V. J. Phys. Chem. A 2010, 114, 8824–8830. (13) Kosenkov, D.; Slipchenko, L. V. J. Phys. Chem. A 2011, 115, 392–401. (14) Thole, B. Chem. Phys. 1981, 59, 341–350. (15) Laumoureux, G.; Roux, B. J. Chem. Phys. 2003, 119, 3025–3039. (16) Rick, S. W.; Stuart, S. J.; Berne, B. J. J. Chem. Phys. 1994, 101, 6141–6156. 20 ACS Paragon Plus Environment

Page 21 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

(17) Rick, S. W.; Stuart, S. J.; Bader, J. S.; Berne, B. J. J. Mol. Liq. 1995, 65–66, 31–40. (18) Rick, S. W.; Berne, B. J. J. Am. Chem. Soc. 1996, 118, 672–679. (19) Lipparini, F.; Barone, V. J. Chem. Theory Comput. 2011, 7, 3711–3724. (20) Lipparini, F.; Cappelli, C.; Barone, V. J. Chem. Theory Comput. 2012, 8, 4153–4165. (21) Lipparini, F.; Cappelli, C.; Scalmani, G.; De Mitri, N.; Barone, V. J. Chem. Theory Comput. 2012, 8, 4270–4278. (22) Miertus, S.; Scrocco, E.; Tomasi, J. Chem. Phys. 1981, 55, 117–129. (23) Cances, E.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 1997, 107, 3032–3041. (24) Mennucci, B.; Cances, E.; Tomasi, J. J. Phys. Chem. B 1997, 101, 10506–10517. (25) Tomasi, J.; Mennucci, B.; Cammi, R. Chem. Rev. 2005, 105, 2999–3093. (26) Cramer, C.; Truhlar, D. Chem. Rev. 1999, 99, 2161–2200. (27) Cossi, M.; Scalmani, G.; Rega, N.; Barone, V. J. Chem. Phys. 2002, 117, 43–54. (28) Curutchet, C.; Muñoz-Losa, A.; Monti, S.; Kongsted, J.; Scholes, G. D.; Mennucci, B. J. Chem. Theory Comput. 2009, 5, 1838–1848. (29) Curutchet, C.; Bidon-Chanal, A.; Soteras, I.; Orozco, M.; Luque, F. J. Phys. Chem. B 2005, 109, 3565–3574. (30) Soteras, I.; Curutchet, C.; Bidon-Chanal, A.; Orozco, M.; Luque, F. J. Mol. Struc.-Theochem 2005, 727, 29–40. (31) Lunkenheimer, B.; Köhn, A. J. Chem. Theory Comput. 2013, 9, 977–994. (32) Schwabe, T.; Sneskov, K.; Haugaard Olsen, J. M.; Kongsted, J.; Christiansen, O.; Hättig, C. J. Chem. Theory Comput. 2012, 8, 3274–3283. 21 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

(33) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G.; Guido, C. A.; Mennucci, B.; Scalmani, G.; Frisch, M. J. Chem. Science 2011, 2, 2143–2161. (34) Scalmani, G.; Frisch, M. J. J. Chem. Phys. 2010, 132, 114110. (35) Lipparini, F.; Scalmani, G.; Mennucci, B.; Cancès, E.; Caricato, M.; Frisch, M. J. J. Chem. Phys. 2010, 133, 014106. (36) Christiansen, O.; Mikkelsen, K. V. J. Chem. Phys. 1999, 110, 1365–1375. (37) Christiansen, O.; Mikkelsen, K. V. J. Chem. Phys. 1999, 110, 8348–8360. (38) Osted, A.; Kongsted, J.; Mikkelsen, K. V.; Christiansen, O. Mol. Phys. 2003, 101, 2055–2071. (39) Kongsted, J.; Pedersen, T. B.; Osted, A.; Hansen, A. E.; Mikkelsen, K. V.; Christiansen, O. J. Phys. Chem. A 2004, 108, 3632–3641. (40) Steindal, A. H.; Ruud, K.; Frediani, L.; Aidas, K.; Kongsted, J. J. Phys. Chem. B 2011, 115, 3027–3037. (41) Caricato, M.; Mennucci, B.; Scalmani, G.; Trucks, G. W.; Frisch, M. J. J. Chem. Phys. 2010, 132, 084102. (42) Cammi, R.; Fukuda, R.; Ehara, M.; Nakatsuji, H. J. Chem. Phys. 2010, 133, 024104. (43) Fukuda, R.; Ehara, M.; Nakatsuji, H.; Cammi, R. J. Chem. Phys. 2011, 134, 104109. (44) Cammi, R. J. Chem. Phys. 2009, 131, 164104. (45) Cammi, R. Int. J. Quantum Chem. 2010, 110, 3040–3052. (46) Caricato, M.; Scalmani, G.; Trucks, G. W.; Frisch, M. J. J. Phys. Chem. Lett. 2010, 1, 2369– 2373. (47) Caricato, M. J. Chem. Phys. 2011, 135, 074113.

22 ACS Paragon Plus Environment

Page 23 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

(48) Caricato, M.; Scalmani, G.; Frisch, M. J. J. Chem. Phys. 2011, 134, 244113. (49) Caricato, M. J. Chem. Theory Comput. 2012, 8, 5081–5091. (50) Caricato, M. J. Chem. Theory Comput. 2012, 8, 4494–4502. (51) Stanton, J. F.; Bartlett, R. J. J. Chem. Phys. 1993, 98, 7029–7039. (52) Stanton, J. F. J. Chem. Phys. 1993, 99, 8840–8847. (53) Stanton, J. F.; Gauss, J. J. Chem. Phys. 1994, 100, 4695–4698. (54) Kallay, M.; Gauss, J. J. Chem. Phys. 2004, 121, 9257–9269. (55) Mortier, W. J.; Van Genechten, K.; Gasteiger, J. J. Am. Chem. Soc. 1985, 107, 829–835. (56) Rappe, A.; Goddard, W. J. Phys. Chem. 1991, 95, 3358–3363. (57) Lipparini, F.; Scalmani, G.; Mennucci, B.; Frisch, M. J. J. Chem. Theory Comput. 2011, 7, 610–617. (58) Klamt, A.; Schüürmann, G. J. Chem. Soc., Perkin Trans. 2 1993, 799–805. (59) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995–2001. (60) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669–681. (61) Scalmani, G.; Barone, V.; Kudin, K.; Pomelli, C.; Scuseria, G.; Frisch, M. Theor. Chem. Acc. 2004, 111, 90–100. (62) Olivares del Valle, F.; Tomasi, J. Chem. Phys. 1991, 150, 139–150. (63) Aguilar, M. A.; Olivares del Valle, F. J.; Tomasi, J. Chem. Phys. 1991, 150, 151–161. (64) Mennucci, B.; Cammi, R.; Tomasi, J. J. Chem. Phys. 1998, 109, 2798–2807. (65) Cammi, R.; Corni, S.; Mennucci, B.; Tomasi, J. J. Chem. Phys. 2005, 122, 104513.

23 ACS Paragon Plus Environment

Journal of Chemical Theory and Computation

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

(66) Dunning, T. H. J. Chem. Phys. 1989, 90, 1007–1023. (67) Barone, V.; Cimino, P.; Stendardo, E. J. Chem. Theory Comput. 2008, 4, 751–764. (68) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. B 2009, 113, 6378–6396. (69) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Parandekar, P. V.; Mayhall, N. J.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian Development Version, Revision H.09+ ed.; Gaussian, Inc.: Wallingford, CT, 2010.

24 ACS Paragon Plus Environment

Page 25 of 25

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Chemical Theory and Computation

Graphical TOC Entry

25 ACS Paragon Plus Environment